Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Article

pubs.acs.org/Macromolecules

Efficient and Heteroselective Heteroscorpionate Rare-Earth-Metal


Zwitterionic Initiators for ROP of rac-Lactide: Role of σ‑Ligand
Zehuai Mou,†,‡ Bo Liu,† Xinli Liu,† Hongyan Xie,†,‡ Weifeng Rong,†,‡ Lei Li,†,‡ Shihui Li,†
and Dongmei Cui*,†

State Key Laboratory of Polymer Physics and Chemistry, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences,
Changchun 130022, People’s Republic of China

University of Chinese Academy of Sciences, Changchun Branch, Changchun 130022, People’s Republic of China
*
S Supporting Information
Downloaded via EAST CHINA UNIV SCIENCE & TECHLGY on June 11, 2023 at 13:08:46 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: A series of oxophosphine (3,5-Me2Pz)2CHP-


(R)2O (Pz = pyrazole; R = tBu (HL1), Cy (HL2)) and
iminophosphine (3,5-Me2Pz)2CHP(R)2NAr (R = Cy, Ar = Ph
(HL3); R = Ph, Ar = Ph (HL4), Ar = 2,6-Me2-phenyl (HL5))
heteroscorpionate ligands were synthesized. Abstraction of the
methide proton of these ligands by rare-earth-metal tris(alkyl)s,
Ln(CH2SiMe3)3(THF)2, afforded the corresponding zwitter-
ionic bis(alkyl) complexes L1−5Ln(CH2SiMe3)2(THF) (L1, Ln
= Y (1a), Lu (1b); L2, Ln = Y (2a), Lu (2b); L3, Ln = Y (3a),
Lu (3b); L4, Ln = Y (4a), Lu (4b); L5, Ln = Y (5a), Lu (5b),
while metathesis reaction of the lithium salts of LiL3 and LiL4
with YCl3(THF)2 or YBr3(THF)2 followed by treatment with
LiCH2SiMe3 and KN(SiHMe2)2, respectively, afforded the first
heteroscorpionate yttrium mixed halogen/alkyl or amido complexes L3−4Y(Cl)(CH2SiMe3)(THF) (L3 (6a), L4 (7a)),
L3−4Y(Cl)(N(SiHMe2)2)(THF) (L3 (8a), L4 (9a)), L4Y(Br)(CH2SiMe3)(THF) (10a), and L4Y(Br)(N(SiHMe2)2)(THF)
(11a). The structures of these complexes were well-defined, and the molecular structures of 1a, 2a, 3b, 4b, 5a, and 7a were
further characterized by single crystal X-ray diffraction analysis. Complexes 1−5 showed similar high activity toward the ROP of
rac-LA at room temperature, and both the alkyl species participated in initiation, of which the lutetium complexes exhibited
slightly higher selectivity than their yttrium analogues (Pr = 0.85−0.89 vs 0.80−0.84) despite the bulkiness of the ligands.
Interestingly, the mixed halogen complexes 6a−11a were single-site initiators, where the σ-halogen moiety remaining on the
central metal showed, for the first time, facilitating the heteroselectivity up to Pr = 0.98. This result sheds new light on designing
specifically selective catalytic precursors.

■ INTRODUCTION
Biodegradable and biocompatible polylactides (PLAs) have
might be attributed to their large ionic radii that require steric
encumbered ligands to compensate the thus coordination
become one of the most promising environmentally benign unsaturation. The amino-bis(phenolate) ligated yttrium com-
materials and widely applied in medical and pharmaceutical plexes invented by Carpentier,11 Mountford,12 and our group13
industries as well as daily used plastics anticipated to partly can catalyze ROP of rac-LA in a living and even immortal
replace polyolefins.1 The versatile applications significantly rely manner with high heterotactic selectivity (Pr > 0.96). Similar
on the variable microstructures of PLAs that are governed by high selectivity had been achieved by Okuda et al. with 1,ω-
the initiators employed via ring-opening polymerization (ROP) dithiaalkanediyl-bridged bis(phenolato) scandium complexes.14
processes, in particular, when using rac-LA as the monomer. To Williams reported the highly active dimeric bis(oxo/
access the controllable stereoregulated PLAs,2 the major thiophosphinic)diamido ligated yttrium complexes15 and the
strategy is to synthesize novel complexes by designing various moderate heteroselective monomeric bis(oxophosphinic)amido
bulky ligands to match the steric demanding of the central and phospha-Salen yttrium complexes by introducing a
metals. For example, high isotactic PLA can be obtained with nitrogen atom on the linking bridge (Pi = 0.84).15e,16,17
some Salen (tetradentate phenoxyimine) aluminum alkoxide or Recently, isotactic-enriched PLAs have been realized with rare-
alkyl complexes,3 while heterotactic PLA can be synthesized by earth-metal complexes by Carpentier,18 Arnold,19 and Otero.20
using Salan (tetradentate phenoxyamine) aluminum initiators,4 In general, the selectivity is governed by the ligand geometry
β-diiminate (BDI) zinc and magnesium alkoxides,5 scorpionate
calcium complexes,6 lithium tert-butoxide,7 and indium,8 etc. Received: January 27, 2014
complexes.9 Rare-earth-metal complexes are known as highly Revised: March 17, 2014
active initiators in this field10 albeit with low selectivity, which Published: March 27, 2014

© 2014 American Chemical Society 2233 dx.doi.org/10.1021/ma500209t | Macromolecules 2014, 47, 2233−2241
Macromolecules Article

Scheme 1. Synthesis of Zwitterionic Rare-Earth-Metal Complexes 1−5

Figure 1. ORTEP diagram of the molecular structures of complexes 1a (left), 4b (middle), and 5a (right). Thermal ellipsoids are drawn at the 35%
probability level. All of the hydrogen atoms are omitted for clarity.

around the metal center via the “enantiomorphic-site compounds (3,5-Me2Pz)2CHPR2 (Pz = pyrazole; R = tBu, Cy,
control”19−21 or “chain-end control”16,17,22 mechanism. To or Ph) were synthesized with a similar method we reported
date, however, no reports have been related to the role of the σ- previously.25 Treatment of these compounds with hydrogen
bonded ligands in controlling the selectivity, which are usually peroxide in dichloromethane or with organoazides in THF gave
considered to participate in the initiation. the targeted oxophosphine heteroscorpionate ligands (3,5-
Herein, we report rare-earth-metal complexes supported by Me2Pz)2CHP(tBu)2O (HL1) and (3,5-Me2Pz)2CHP(Cy)2O
new heteroscorpionate ligands23 bearing various bulky (HL2) and the iminophosphine heteroscorpionate ligands
substituent groups that show high activity and moderate to (3,5-Me 2 Pz) 2 CHP(Cy) 2 NPh (HL 3 ), (3,5-Me 2 Pz) 2 CHP-
high hetereoselectivity toward the ROP of rac-LA. More (Ph)2NPh (HL4), and (3,5-Me2Pz)2CHP(Ph)2NAr(2,6-Me2)
importantly, we demonstrate that introducing an inert σ- (HL5) (Scheme 1).
bonded halogen group to these complexes affords the rare- The ligands HL1−HL5 reacted with rare-earth-metal tris-
earth-metal initiators with excellent hetereoselectivity (Pr = (alkyl)s, Ln(CH2SiMe3)3(THF)2 (Ln = Y, Lu), at −30 °C to
0.98), which is the first example that the σ-bonded ligand give complexes 1−4 and 5a, while at 50 °C to afford complex
facilitates the selectivity, although the “halogen determines 5b (Scheme 1). Complexes 1−4 were moisture and air sensitive
regularity” had been accepted long time ago when using but stable at room temperature under a nitrogen atmosphere,
Ziegler−Natta catalysts for propylene and conjugated diene while complexes 5 gradually changed from colorless to yellow
polymerizations.24 in solution. The 1H NMR spectra of complexes 1−5 revealed

■ RESULTS AND DISCUSSION


Synthesis and Characterization of Complexes 1−5.
that the resonances arising from the methine protons of the
ligands disappeared, suggesting the formation of carbanions.26
There are one set of signals for the two metal−alkyl moieties
The phosphine modified bis(3,5-dimethylpyrazolyl)methane and singlet resonances for the pyrazole protons of 4-H, 3-Me,
2234 dx.doi.org/10.1021/ma500209t | Macromolecules 2014, 47, 2233−2241
Macromolecules Article

Scheme 2. Synthesis of Zwitterionic Yttrium Mixed Halogen/Alkyl (Amido) Complexes 6a−11a

and 5-Me, indicating that the two alkyl groups and the two 5.40, and 5.43 ppm, respectively, which shift downfield as
pyrazole rings are equivalent, respectively. The molecular compared to that in homoleptic complex {Y[N-
structures of 1−5 in solid states were defined by single crystal (SiHMe2)2]3(THF)} (4.94 ppm),28 suggesting the absence of
X-ray diffraction analysis (Figure 1, 1a (left), 4b (middle), and β(Si−H) agnostic interaction11e observed frequently in the
5a (right); for 2a and 3b see the Supporting Information Figure bis(dimethylsilyl)amido rare-earth complexes.20,22a,29 Note-
S33). In general, the metal ion is capped by the tridentate worthy was that the two pyrazole rings are equivalent although
NNO or NNN ligand in a κ3-coordination fashion without there are no symmetric centers in molecule structures, owing to
direct bond linkage to the apical carbanion and coordinated to a the probable rapid exchange positions between the halogen
solvated THF and two alkyl groups, adopting a distorted ligand and the silylalkyl (or silylamido) group. These solution-
tetragonal bipyramidal geometry close to the previously state structures were consistent with that established by X-ray
reported complex (3,5-Me2Pz)2CP(Ph)2OY- diffraction as shown in Figure 2 (7a). The heteroscorpionate
(CH2SiMe3)2(THF).25 For complexes 1a and 2a, the bond moiety adopts the same coordination mode to that in complex
lengths of Y(1)−O(1) between the yttrium ion and the oxygen 4b. C(11), P(1), N(5), Y(1), and O(1) generated a plane
from the ligand, 2.207(3) Å (1a) and 2.218(2) Å (2a) are
consistent with those in the literature,25,27 but significantly
shorter than those of Y(1) −O(2) between the yttrium ion and
the oxygen from THF, 2.405(3) Å (1a) and 2.404(2) Å (2a),
suggesting that the coordination of THF molecule is weak and
fluxional. For complexes 3b and 4b, the bond lengths of
lutetium and phosphine imodo nitrogen Lu(1)−N(5),
2.298(3) Å (3b) and 2.341(3) Å (4b) are much shorter than
those between lutetium to the pyrozolyl nitrogen atoms
(average 2.491 Å for 3b and 2.487 Å for 4b). Variation of
the substituents of the ligand framework on the phosphorus
atom affects the orientation of the phenyl ring on N(5). The
phenyl ring is almost parallel to planar NPCO framework in 4b,
which deviates from the corresponding plane in 3b but
perpendicular to the corresponding plane in 5a owing to the
steric repulsion.
Synthesis and Characterization of Monohalogen
Complexes 6a−11a. Ligands HL3 and HL4 were lithiated
by RLi via alkane elimination, which were followed sequentially
by metathesis reaction with yttrium chloride YCl3(THF)2 and
alkylation with LiCH 2 SiMe 3 or amination with K[N-
(SiHMe2)2], to give the corresponding mixed chloride/alkyl
complexes 6a and 7a and the mixed chloride/amido complexes
8a and 9a, respectively. The yttrium mixed bromide/alkyl and Figure 2. ORTEP diagram of the molecular structure of complex 7a.
Thermal ellipsoids are drawn at the 35% probability level. Hydrogen
amido complexes 10a and 11a were prepared in the same
atoms have been omitted for clarity. Selected bond lengths (Å) and
procedure by using YBr3(THF)2 instead (Scheme 2). In the 1H angles (deg): N(1)−Y(1) 2.452(3), N(3)−Y(1) 2.565(3), N(5)−Y(1)
NMR spectra of these complexes, the integral intensity ratio of 2.377(2), O(1)−Y(1) 2.380(2), C(30)−Y(1) 2.408(3), Cl(1)−Y(1)
the silylalkyl (or silylamido) group to the ligand is 1:1. The 2.5632(10); N(5)−Y(1)−N(1) 88.36(9), N(5)−Y(1)−N(3)
silylhydride SiHMe2 in 8a, 9a, and 11a gives a septet at 5.19, 91.76(8), C(30)−Y(1)−Cl(1) 104.81(9).

2235 dx.doi.org/10.1021/ma500209t | Macromolecules 2014, 47, 2233−2241


Macromolecules Article

Table 1. ROP of rac-LA Initiated by Heteroscorpionate Rare-Earth Bis(alkyl) Complexesa


run cat. T (°C) t (min) convb (%) Mn,calcdc (104) Mn,expd (104) Mw/Mnd Pre
1 1a 25 15 98 1.41 1.88 1.48 0.82
2 1b 25 15 97 1.40 3.10 1.63 0.87
3 2a 25 15 99 1.43 1.70 1.80 0.83
4 2b 25 15 97 1.40 2.94 2.05 0.86
5 1a 5 10 94 1.35 2.19 1.72 0.86
6 1b 5 30 42 0.60 2.69 1.72 0.91
7 2a 5 10 97 1.40 3.12 1.95 0.85
8 2b 5 30 88 1.27 2.60 2.12 0.89
9 3a 25 15 97 1.40 3.56 1.15 0.84
10 3b 25 15 98 1.41 2.78 1.13 0.89
11 4a 25 15 98 1.41 3.67 1.20 0.84
12 4b 25 15 97 1.40 2.43 1.12 0.85
13 5a 25 15 97 1.41 2.85 1.26 0.80
14 5b 25 15 96 1.38 2.63 1.13 0.83
15f 1b 25 10 98 2.82 3.60 2.07 0.78
16g 1b 25 10 85 2.50 3.34 1.84 0.88
17f 3b 25 10 99 2.85 3.68 1.49 0.83
18g 3b 25 10 57 1.64 2.31 1.19 0.89
a
Conditions: [rac-LA]0 = 0.8 M, solvent (THF), [M]/cat. = 200. bDetermined by 1H NMR spectrum. cMn,calcd = [M]/2cat. × 144 × conv (%).
d
Determined by GPC against polystyrene standard, Mn using a correcting factor for polylactides (0.58). eDetermined by homonuclear decoupled 1H
NMR spectrum. fSolvent (5 mL of toluene). [M]/cat. = 400. g[M]/cat. = 400.

adopting the meridional configuration that bisects the angle catalytic performances of complexes 3−5 were more promising
formed by the chlorine and the silylalkyl groups. The Cl(1)− as compared to our previously reported complexes supported
Y(1) bond length (2.5632 Å) is slightly shorter than the by the iminophosphine hetereoscorpionate with a methoxy side
yttrium terminal chloride bond lengths reported in the group (3,5-Me2Pz)2CP(Ph)2N(C6H4-o-OMe)Ln(CH2SiMe3)2
literature,30 owing to the enhanced Lewis acidity of the metal (conversion 98% in 15 min, Pr = 0.80−0.89 vs conversion
center. 95% in 5 h, Pr = 0.68−0.7025), as the solvated THF molecule
Ring-Opening Polymerization of rac-LA Initiated by coordinates to the metal center is flexible and reversible in
Bis(alkyl) Complexes. All the zwitterionic rare-earth-metal complexes 3−5 that facilitates to improve the stereoselectivity
bis(alkyl) complexes 1−5 were employed as single-component of the catalytic systems,15e while in the late case the
catalysts to initiate the ROP of rac-LA at room temperature to coordination of the methoxy group to the metal center is
reach conversions over 94% within 15 min with hetero- strong and irreversible. Whereas, THF was not the optimum
selectivity Pr ranging from 0.80 to 0.89 depending on the medium for the current polymerization systems, given as Table
electronics and sterics of ligands as well as central metals (Table 1 runs 15−18, the polymerization with 3b proceeded faster in
1, runs 1−4 and 9−14). When the polymerization was toluene to transfer 400 equiv of rac-LA to PLA in 10 min while
conducted at a lower temperature (5 °C), the yttrium only 57% conversion could be reached in THF under the same
complexes 1a and 2a maintained superior activity (up to 94% conditions, although the stereoregularity of the resulting PLAs
conversion in 10 min) as compared with the lutetium decreased slightly. This was ascribed to the competition
counterparts 1b and 2b, while the polymerization initiated by between THF and rac-LA for coordinating to the active
complexes bearing dicyclohexylphosphine substituents pro- metal centers (toluene usually does not coordinate to metal
ceeded faster than that initiated by the corresponding analogues center when a polar monomer polymerization performed in
bearing di-tert-butylphosphine substituents (2a vs 1a and 2b vs it).15a,b,e,22b Noteworthy was that complexes 4 showed a slightly
1b) (Table 1, runs 5−8). In contrast, all the lutetium complexes higher heteroselectivity than the more steric bulky complexes 5
1b and 2b showed higher hetereoselectivity than the yttrium (Table 1, runs 13 and 14), which might be explained by the
counterparts 1a and 2a despite the steric bulkiness of the vertical positioned phenyl rings in complexes 4 (the phenyl ring
ligands and polymerization temperature (Table 1, runs 1−8). takes perpendicular position to the ligand framework, leaving a
The oxophosphine heteroscorpionate complexes 1 and 2 were more opening sphere to the metal center) that orientates the
comparable to the analogous (3,5-Me2Pz)2CP(Ph)2OY- monomer coordination−insertion during the propagation
(CH2SiMe3)2(THF) reported previously by us in the aspect cycle.6a,11b,22b,31
of stereoselectivity but provided much higher catalytic activity Generally, monoanionic auxiliary ligand supported rare-
(conversion 97−99% in 15 min vs conversion 95% in 1 h25). earth-metal bis(alkyl) complexes are double-site initiators
The iminophosphine hetereoscorpionate ligated complexes 3, because of the highly active nature of the metal−carbon
4, and 5 exhibited similar performances as complexes 1 and 2 bonds, although a few bis(alkyl) (or bis(amido)) complexes
but provided PLAs with controlled molecular weight and have been found of single-site where the two alkyl (or amido)
narrow molecular weight distribution (Table 1, runs 9−14), groups take endo (less active) or exo (active) positions against
which might be contributed to the more steric bulkiness of the the ligands.15b,32 The polymerization data listed in Table 1
iminophosphine hetereoscorpionate ligands than the more show that the molecular weights measured by GPC for most of
opening coordination sphere around metal center generated by the resulting PLAs were larger than the theoretical values
the oxophosphine hetereoscorpionate ligands. In addition, the calculated based on both alkyls participating in the initiation.
2236 dx.doi.org/10.1021/ma500209t | Macromolecules 2014, 47, 2233−2241
Macromolecules Article

Thus, complex 3b was chosen to ascertain single or double sites Ring-Opening Polymerization of rac-LA Initiated by
of the current systems (Table 2). At the lower monomer-to- Monohalogen Complexes. These novel mixed halogen/alkyl
and halogen/amido complexes 6a−11a were subjected to
Table 2. Effect of Monomer-to-Initiator Ratioa initiate ROP of rac-LA, which showed lower activity by taking 1
h to transfer 200 equiv of monomers into polymers compared
convb Mn,calcdc Mn,calcdd Mn,expe
run [M]/[I] (%) (104) (104) (104) Mw/Mne with 15 min for the bis(alkyl) analogues 3a and 4a. This might
be ascribed to the single active site per metal against the double
1 100 94 0.68 1.35 1.30 1.21
sites in the later two cases, as the strong σ-bonded chlorine/
2 200 96 1.38 2.76 2.15 1.25
bromine did not participated in the initiation, and the presence
3 400 94 2.71 5.41 3.09 1.39
of which retarded the monomer insertion into the active site via
4 600 89 3.84 7.69 4.53 1.27
increasing the steric bulkiness around the metal center. Thus,
5 800 98 5.64 11.3 5.16 1.54
the bromido complexes 10a and 11a exhibited slightly lower
6 1000 91 6.55 13.1 7.81 1.26
activity than their chloride analogues 7a and 9a (Table 3, runs
a
Conditions: 10 μmol of initiator (3b), solvent (THF), T = 25 °C, 8, 9 vs runs 3, 6), respectively. The molecular weight of the
reaction time (2 h) was not optimized. bDetermined by 1H NMR resultant PLA was in consistence with the theoretical value
spectrum. cMn,calcd = [M]/2[I] × 144 × conv (%). dMn,calcd = [M]/[I]
× 144 × conv (%). eDetermined by GPC against polystyrene standard,
calculated following the single-site active species. Strikingly,
Mn using a correcting factor for polylactides (0.58). however, this σ-bonded halogen group endowed the metal
center excellent heteroselectivity up to Pr = 0.98 (Figure 4), no
initiator ratios 100:1−200:1, the molecular weight of the matter whether the other σ-bonded initiator was alkyl or amido
resulting PLA was close to the theoretic value calculated as the (Table 3, runs 1−6, 8, and 9), which was much higher than the
single-site initiator, while with the ratio increasing from 400:1 bis(alkyl) complexes 3−4 bearing the same ligands and
to 1000:1, the molecular weight was comparable to that arising competitive with the best records in the literature.9c,13a In
from the double-site initiators. Thus, we suggested in this addition, the distinguished selectivity was maintained even at
system that both alkyl groups participated in the initiation, and elevated polymerization temperatures (9a, Pr = 0.90, Tp = 60
the deviation of the experimental values under the lower °C) (Table 3, run 7). This suggested that the σ-halogen group
monomer-to-initiator ratios from the theoretic ones was remained on the active center to orient the monomer
ascribed to the aggregation of the active species under higher coordination and insertion. The 1H NMR and MALDI-TOF
concentrations.33 analyses of an L-LA oligomer obtained with complex 9a
The polymerization kinetic behaviors for complexes 3a, 4a, revealed that the molecule chain is clearly capped with
5a, and 3b in THF at room temperature were further −N(SiHMe2)2 group and hydroxyl group at both ends (Figures
investigated ([LA]0 = 0.5 M, [LA]/[cat.] = 300). The S35 and S36), suggesting that the ROP of rac-LA in this system
polymerization data showed linear fits to plots of ln{[M]0/ was indeed conducted via the coordination insertion mecha-
[M]t} vs time (Figure 3) to establish the first-order kinetics in nism. The thus “halogen-control” selectivity has been accepted
with respect to the iso-specifically selective propylene polymer-
ization and cis-1,4 regioselective polymerization of conjugated
dienes by using Ziegler−Natta catalysts,24a−d which, regarding
the polymerization of polar monomers, has not been reported
yet.

■ CONCLUSION
We report a series of new bis(alkyl) rare-earth-metal
zwitterionic complexes supported by oxophosphine and
iminophosphine heteroscorpionate ligands that act as double-
site initiators for the polymerization of rac-LA to exhibit
significantly enhanced activity and high heteroselectivity.
Polymerization kinetic study suggests these systems present
the first-order on the monomer concentration, among which
the yttrium-based initiators are more active, while the lutetium-
based ones are highly specific selective. Strikingly, the mixed
halogen/alkyl and amido complexes bearing the similar
heteroscopionate ligands display as single-site initiators, as the
Figure 3. Semilogarithmic plots of rac-LA conversion versus time (T = σ-bonded halogen group does not participate in the initiation
25 °C, in THF, [LA]0 = 0.5 M, [LA]/[cat.] = 300): 3a (black squares, that facilitates to the distinguished heteroselectivity of Pr = 0.98
kapp = 4.59 × 10−3 s−1, R2 = 0.993), 4a (red circles, kapp = 2.96 × 10−3
s−1, R2 = 0.992), 5a (blue triangles, kapp = 2.49 × 10−3 s−1, R2 = 0.999),
at room temperature. This work shows for the first time the σ-
3b (purple stars, kapp = 1.07 × 10−3 s−1, R2 = 0.998). bonded ligand controls the selectivity more powerfully than
changing the substituents and metal centers, which sheds new
light on designing specific-selective initiators.
monomer concentration in each case. The slopes clearly
indicated the influences of ligands on polymerization rates in an
order of 3a > 4a > 5a and a much faster rate for the yttrium
■ EXPERIMENTAL SECTION
General Procedures. All reactions were carried out under a dry
complex than its lutetium analogue (kapp = 4.59 × 10−3 s−1 for nitrogen atmosphere using standard Schlenk techniques or in a
3a vs 1.07 × 10−3 s−1 for 3b), which is comparable to that glovebox filled with dry nitrogen. Hexane and toluene were purified
reported by Williams.15e using an SPS Braun system. THF was dried by distillation over sodium

2237 dx.doi.org/10.1021/ma500209t | Macromolecules 2014, 47, 2233−2241


Macromolecules Article

Table 3. ROP of rac-LA Initiated by Rare-Earth Mono(alkyl/amido) Complexesa


run cat. M/cat. T (°C) t (h) convb (%) Mn,calcdc (104) Mn,expd (104) Mw/Mnd Pre
1 6a 200 25 1 95 2.74 2.38 1.10 0.93
2 6a 200 0 7 97 2.79 2.96 1.14 0.96
3 7a 200 25 1 94 2.71 3.02 1.20 0.97
4 7a 200 0 10 97 2.79 3.59 1.37 0.98
5 8a 200 25 2 86 2.48 3.17 1.30 0.93
6 9a 200 25 1 88 2.53 3.93 1.28 0.98
7 9a 400 60 0.5 90 5.18 6.68 1.35 0.90
8 10a 200 25 1 87 2.51 2.99 1.31 0.98
9 11a 200 25 1 77 2.21 2.87 1.39 0.98
a
Conditions: [rac-LA]0 = 0.8 M, solvent (THF). bDetermined by the 1H NMR spectrum. cMn,calcd = [M]/cat. × 144 × conv (%). dDetermined by
GPC against polystyrene standard, Mn using a correcting factor for polylactides (0.58). eDetermined by the homonuclear decoupled 1H NMR
spectrum.

follows: A solution of complex 1a (14.0 mg, 20 μmol, [LA]/[Y] =


200:1) in THF (2 mL) was added to a stirred solution of rac-LA
(0.576 g, 4 mmol) in THF (3 mL). The polymerization took place
immediately at room temperature. The system became viscous in a few
minutes. It was stirred for 15 min and then quenched by adding several
drops acidified ethanol, and a small sample of the viscous solution was
separated for the measurement of conversion by 1H NMR. Then
polymers were precipitated with abundant ethanol, collected, and dried
at 40 °C for 24 h in vacuo. The molecular weight and polydispersity
index of the resulting polymers were determined by GPC. The tacticity
of the PLA was calculated according to the methine region
homonuclear decoupled 1H NMR spectrum.
Synthesis of Proligands and Representative Complexes.
Synthesis of (3,5-Me 2 Pz) 2 CHP( t Bu) 2 . A solution of bis(3,5-
dimethylpyzazoly)methane (4.08 g, 20 mmol) in THF in a Schlenk
flask was cooled down to −65 °C, followed by slow addition of nBuLi
in hexane (12.5 mL, 1.65 mol/L) with a syringe over 25 min under
vigorous agitation, which immediately resulted in a white suspension.
Figure 4. Methine region of the homonuclear decoupled 1H NMR The mixture was stirred for another 2 h at this temperature. Then a
spectra (400 MHz, CDCl3, 298 K) of the PLAs obtained from 4a solution of tBu2PCl (3.61 g, 20 mmol) in 10 mL of THF was added
(curve 2, run 11 in Table 1) and 9a (curve 1, run 6 in Table 3) at dropwise over 20 min, which generated a gray-green solution. The
room temperature. reaction mixture was heated to 50 °C and stirred for an additional 12
h, which turned into dark red solution. Removal of the THF led to
with benzophenone as indicator under a nitrogen atmosphere and was brown solids, and toluene (150 mL) was added to extract. The
stored over freshly cut sodium in a glovebox. Rare-earth-metal insoluble solid was removed by filtration, and the filtrate was
tris(alkyl)s were synthesized according to a previous report.34 Bis(3,5- evaporated to give an orange solid (yield: 87%). 1H NMR (400
dimethylpyrazolyl)methane was synthesized according to the literature MHz, C6D6): δ 7.12 (s, 1H; C−H), 5.57 (s, 2H; Pz−H), 2.70 (s, 6H;
from 3,5-dimethylpyrazole and dichloromethane catalyzed by Pz−CH3), 2.13 (s, 6H; Pz−CH3), 1.07 (s, 18H; tBu−H), 1.05 ppm (s,
n
Bu4NBr.35 rac-LA was recrystallized with dry ethyl acetate three 18H; tBu−H). 31P NMR (162 MHz, C6D6): δ 44.14 ppm. Anal. Calcd
times. Glassware and flasks using in the polymerization were dried in for C19H33N4P: C, 65.49; H, 9.55; N, 16.08. Found: C, 65.41; H, 9.50;
an oven at 115 °C overnight and exposed to a vacuum−nitrogen cycle N, 16.03.
three times. The molecular weight and molecular weight distribution Synthesis of (3,5-Me2Pz)2CHP(Cy)2. The compound was synthe-
of the polymers were measured using a TOSOH HLC 8220 GPC sized with the same method described as above using chlorodicyclo-
instrument at 40 °C with THF as eluent against polystyrene standards. hexylphosphine (yield: 76%). 1H NMR (400 MHz, CDCl3): δ 6.72 (s,
Organometallic samples for NMR spectroscopic analysis were 1H; C−H), 5.76 (s, 2H; Pz−H), 2.36 (s, 6H; Pz−CH3), 2.18 (s, 6H;
prepared in a glovebox by the use of NMR spectroscopy tubes and Pz−CH3), 1.82−1.80 (m, 2H; Cy−H), 1.71−1.64 (m, 4H; Cy−H),
then sealed by paraffin film. 1H, 31P, and 13C NMR spectra were 1.59−1.58 (m, 2H; Cy−H), 1.46−1.40 (m, 4H; Cy−H), 1.16−1.01
recorded using a Bruker AV400 spectrometer. The MALDI-TOF mass (m, 8H; Cy−H), 0.89−0.83 ppm (m, 2H; Cy−H). 31P NMR (162
spectrum was obtained with a Bruker Daltonic MicroFlex LT at the MHz, C6D6): δ 14.60 ppm. Anal. Calcd for C23H37N4P: C, 68.97; H,
National Analytic Research Center of the Changchun Institute of 9.31; N, 13.99. Found: C, 68.92; H, 9.28; N, 13.65.
Applied Chemistry (CIAC). Synthesis of (3,5-Me2Pz)2CHP(tBu)2O (HL1). Hydrogen peroxide
X-ray Crystallographic Studies. Crystals for X-ray analysis were (1.6 g of a 30% w/w solution, 14 mmol) was added dropwise to a
obtained as described in the following preparations. The crystals were stirred solution of (3,5-Me2Pz)2CHP(tBu)2 (2.44 g, 7 mmol) in
manipulated in a glovebox. Data collection was performed at −86.5 °C CH2Cl2 at 0 °C. The mixture was then warmed to room temperature
using a Bruker SMART APEX diffractometer with a CCD area and stirred for 2 h. Sodium thiosulfate (4.74 g, 30 mmol) aqueous
detector and graphite monochromated Mo Kα radiation (λ = 0.710 73 solution was added to the mixture slowly and stirred for another 1 h.
Å). The determination of crystal class and unit-cell parameters was Separated the aqueous and organic layers, extracted the aqueous layer
carried out by the SMART program package. The raw frame data were with CH2Cl2 twice (2 × 20 mL), collected the organic phase, and dried
processed using SAINT and SADABS to yield the reflection data file. with anhydrous MgSO4 overnight. Removing the solvent under
The structures were solved by using the SHELXTL program. reduced pressure gave a yellow residue. Washed the residue with
ROP of rac-LA. Polymerizations of rac-LA were carried out in a 15 hexane gave HL1 as a white solid (yield: 61%). 1H NMR (400 MHz,
mL flask under a N2 atmosphere. A typical procedure was described as C6D6): δ 6.63 (d, J = 12 Hz, 1H; C−H), 5.62 (s, 2H; Pz−H), 2.23 (s,

2238 dx.doi.org/10.1021/ma500209t | Macromolecules 2014, 47, 2233−2241


Macromolecules Article

6H; Pz−CH3), 2.08 (s, 6H; Pz−CH3), 1.32 (s, 18H; tBu−H), 1.29 reduced pressure, and the residue was extracted with toluene. The
ppm (s, 18H; tBu−H). 31P NMR (162 MHz, C6D6): δ 79.63 ppm. insoluble solid was removed by filtration. The filtrate was concentrated
Anal. Calcd for C19H33N4OP: C, 62.61; H, 9.13; N, 15.37. Found: C, to 2 mL and was added several drops of hexane. Complex 6a was
62.57; H, 9.09; N, 15.33. obtained as pale yellow solids within 2 days when the solution was
Synthesis of (3,5-Me2Pz)2CHP(Cy)2O (HL2). HL2 was synthesized kept at −30 °C. Yield: 53%. 1H NMR (400 MHz, C6D6): δ 7.72 (d, J =
using the same method as HL1 (yield: 65%). 1H NMR (400 MHz, 8 Hz, 2H; Ph−H), 7.32 (t, J = 8 Hz, 2H; Ph−H), 7.01 (t, J = 8 Hz,
CDCl3): δ 6.37 (d, J = 8 Hz, 1H; C−H), 5.83 (s, 2H; Pz−H), 2.40− 1H; Ph−H), 5.65 (s, 2H; Pz−H), 3.63 (s, 4H; THF), 2.37 (s, 6H; Pz−
2.29 (m, 2H; Cy−H), 2.20 (s, 6H; Pz−CH3), 2.18 (s, 6H; Pz−CH3), CH3), 2.22 (s, 6H; Pz−CH3), 1.97, 1.60, 1.48, 1.24, 0.96 (m, 22H;
2.06−0.83 ppm (m, 20H; Cy−H). 31P NMR (162 MHz, C6D6): δ Cy−H), 1.32 (s, 4H; THF), 0.35 (s, 9H; Si−Me3), −0.23 ppm (s, 2H;
52.99 ppm. Anal. Calcd for C23H37N4OP: C, 66.32; H, 8.95; N, 13.45. Y−CH2). 31P NMR (162 MHz, C6D6): δ 39.24 ppm. Anal. Calcd (%)
Found: C, 66.27; H, 8.90; N, 13.40. for C37H60ClN5OPSiY: C, 57.39; H, 7.81; N, 9.04. Found: C, 57.33;
Synthesis of (3,5-Me2Pz)2CHP(Cy)2NPh (HL3). A solution of phenyl H, 7.79; N, 8.89.
azide (0.54 g, 4.5 mmol) in THF (10 mL) was added dropwise to a Synthesis of Complex 8a. To a solution of HL3 (0.3932 g, 0.8
stirred solution of (3,5-Me2Pz)2CHP(Cy)2 (1.2 g, 3.0 mmol) in THF mmol) in THF at −30 °C, a THF solution of LiCH2SiMe3 (0.0753 g,
(40 mL) at room temperature in a glovebox. After 12 h, the solvent 0.8 mmol) was added dropwise. After 30 min, the mixture was added
was removed in vacuo to give a light yellow solid, which was washed to a suspension of YCl3 (0.1562 g, 0.8 mmol) in THF, which was
with hexane three times and isolated by filtration, followed by drying prepared 1 day in advance. With the reaction proceeding, the above
under reduced pressure at room temperature (yield: 65%). 1H NMR reaction suspension became clear gradually. After 7 h, the mixture was
(400 MHz, C6D6): δ 7.29 (t, J = 8 Hz, 2H; Ph−H), 7.10 (d, J = 8 Hz, cooled to −30 °C and a solution of K[N(SiHMe2)2] (0.1303 g, 0.76
2H; Ph−H), 6.87 (t, J = 8 Hz, 1H; Ph−H), 6.63 (d, J = 8 Hz, 1H; Ph− mmol) was added slowly. Then, the reaction mixture was kept stirring
H), 5.60 (s, 2H; Pz−H), 3.04−2.94 (m, 2H; Cy−H), 2.30−2.26 (m, for 5 h at room temperature. The solvent was removed under reduced
2H; Cy−H), 2.12 (s, 6H; Pz−CH3), 1.98 (s, 6H; Pz−CH3), 1.83−0.82 pressure, and the residue was extracted with toluene. The insoluble
ppm (m, 18H; Cy−H). 31P NMR (162 MHz, C6D6): δ 33.34 ppm. solid was removed by filtration. The filtrate was concentrated to 2 mL
Anal. Calcd for C29H42N5P: C, 70.85; H, 8.61; N, 14.24. Found: C, and was added several drops of hexane. Complex 8a was obtained as
70.51; H, 8.50; N, 14.11. white solids within 2 days when the solution was kept at −30 °C.
Synthesis of (3,5-Me2Pz)2CHP(Ph)2NPh (HL4). To a solution of Yield: 60%. 1H NMR (400 MHz, C6D6): δ 7.78 (d, J = 8 Hz, 2H; Ph−
(3,5-Me2Pz)2CHP(Ph)2 (3.88 g, 10 mmol) in THF, phenyl azide H), 7.30 (t, J = 8 Hz, 2H; Ph−H), 7.95 (t, J = 8 Hz, 1H; Ph−H), 5.56
(1.67 g, 14 mmol) diluted with 5 mL of THF was added dropwise in a (s, 2H; Pz−H), 5.19 (sept, J = 3.2 Hz, 2H; Si−H), 3.62 (m, 4H; THF),
glovebox under stirring, and then the yellow solution was heated to 70 2.44 (s, 6H; Pz−CH3), 1.96, 1.78, 1.63, 1.45, 1.13, 0.97, 0.59 (m, 22H;
°C under a nitrogen atmosphere for 2 days. The solvent was removed Cy−H), 2.15 (s, 6H; Pz−CH3), 1.36 (m, 4H; THF), 0.44 ppm (d, J =
in vacuo to give a light yellow solid, which was washed with hexane 3.2 Hz, 12H; SiH−Me2). 31P NMR (162 MHz, C6D6): δ 38.31 ppm.
and isolated by filtration as a white solid, followed by drying under
13
C NMR (100 MHz, C6D6): δ 149.86, 148.34, 148.24 (C3 or C5),
reduced pressure at room temperature (yield: 79%).1H NMR (400 145.11, 145.05, 129.43, 129.42, 122.57, 122.55 (Ph) 106.13 (C4), 69.14
MHz, C6D6), δ 7.96 (m, 4H; Ph−H), 7.19 (m 4H; Ph−H), 7.02 (m, (THF), 54.21 (d, J = 127 Hz; P−C), 37.91, 37.23, 27.83, 27.71, 27.28,
7H; Ph−H), 6.84 (s, 1H; C−H), 5.53 (s, 2H; Pz−H), 1.96 (s, 6H; 27.16, 26.68, 25.81, 25.78, 25.61 (Cy), 26.56 (THF), 14.88 (Pz−CH3),
Pz−CH3), 1.92 ppm (s, 6H; Pz−CH3). Anal. Calcd for C29H30N5P: C, 12.21 (Pz−CH3), 3.52 ppm (SiH−Me2). Anal. Calcd (%) for
72.63; H, 6.31; N, 14.60. Found: C, 72.43; H, 6.18; N, 14.51. C37H63ClN6OPSi2Y: C, 54.23; H, 7.75; N, 10.26. Found: C, 54.16;
Synthesis of (3,5-Me2Pz)2CHP(Ph)2NAr(2,6-diMe) (HL5). HL5 was H, 7.70; N, 10.20.
synthesized using the same method as HL4 (yield: 69%). 1H NMR CCDC-921207 (1a), 921208 (2a), 957722 (3b), 957723 (4b),
(400 MHz, C6D6), δ 8.20−8.14 (m, 4H; Ph−H), 7.07 (m, 8H; Ph− 957724 (5a), and 965061 (7a) contain the supplementary crystallo-
H); 6.81 (m, 1H; Ph−H), 6.72 (d, J = 12.0 Hz, 1H; C−H), 5.45 (s, graphic data for this paper. These data can be obtained free of charge
2H; Pz−H), 2.15 (s, 6H; Ph−CH3), 1.97 (s, 6H; Pz−CH3), 1.55 ppm from The Cambridge Crystallographic Data Centre via www.ccdc.cam.
(s, 6H; Pz−CH3). Anal. Calcd for C31H34N5P: C, 73.35; H, 6.75; N, ac.uk/data_request/cif.


13.80. Found: C, 73.27; H, 6.61; N, 13.59.
Synthesis of Complex 1a. Tris(alkyl)s yttrium complex [Y- ASSOCIATED CONTENT
(CH2SiMe3)3(THF)2] (0.5 mmol) was dissolved in dry hexane and
cooled to −30 °C. A THF suspension of HL1 at −30 °C was added *
S Supporting Information
into the above solution. After stirring for 4 h at room temperature, the Experimental details for complexes 1b, 2−5, 7a, and 9a−11a;
reaction mixture was concentrated to a small portion and was added 1
H and 31P NMR spectra of all complexes; table giving selected
several drops of hexane. Colorless crystal deposited at the bottom of bond lengths and bond angles; molecular structures of
the bottle from the solution under −30 °C after 3 days. Yield: 75%. 1H
NMR (400 MHz, C6D6): δ 5.56 (s, 2H; Pz−H), 3.58 (s, 4H; THF),
complexes 2a and 3b; the crystallographic data for complexes
2.19 (s, 6H; Pz−CH3), 2.14 (s, 6H; Pz−CH3), 1.19 (s, 4H; THF), 1.04 1a, 2a, 3b, 4b, 5a, and 7a; 1H NMR and MALDI-TOF mass
(s, 9H; tBu−H), 1.01 (s, 9H; tBu−H), 0.60 (s, 18H; Si−Me3), −0.28 spectra of an oligomer. This material is available free of charge
ppm (s, 4H; Y−CH2). 31P NMR (162 MHz, C6D6): δ 75.22 ppm. 13C via the Internet at http://pubs.acs.org.


NMR (100 MHz, C6D6): δ 148.29 (C3 or C5), 145.06, 144.99 (C3 or
C5), 105.76 (C4), 70.50 (THF), 63.92 (d, J = 67 Hz; P−C), 36.32 (P− AUTHOR INFORMATION
C−Me3), 35.59 (P−C−Me3), 31.08 (Y−CH2), 30.71 (Y−CH2), 28.36
(P−CMe3), 25.18 (THF), 14.54 (Pz−CH3), 12.74 (Pz−CH3), 5.08 Corresponding Author
ppm (Si−Me3). Anal. Calcd for C31H62N4O2PSi2Y: C, 53.27; H, 8.94; *E-mail dmcui@ciac.ac.cn; Fax (+86) 431 85262774; Tel +86
N, 8.02. Found: C, 53.20; H, 8.65; N, 7.89. 431 85262773 (D.C.).
Synthesis of Complex 6a. To a solution of HL3 (0.3932 g, 0.8
mmol) in THF at −30 °C, a THF solution of LiCH2SiMe3 (0.0753 g, Notes
0.8 mmol) was dropwise added under stirring. After 30 min, the The authors declare no competing financial interest.


mixture was added to a suspension of YCl3 (0.1562 g, 0.8 mmol) in
THF, which was prepared 1 day in advance. With the reaction ACKNOWLEDGMENTS
proceeding, the above reaction suspension became clear gradually.
After 7 h, the mixture was cooled to −30 °C and a solution of This work was partially supported by The National Natural
LiCH2SiMe3 (0.0716 g, 0.76 mmol) was added slowly, then being Science Foundation of China for project nos. 51321062,
stirred for 3 h at room temperature. The solvent was removed under 21274143, and 21361140371.
2239 dx.doi.org/10.1021/ma500209t | Macromolecules 2014, 47, 2233−2241
Macromolecules


Article

REFERENCES (12) Clark, L.; Cushion, M. G.; Dyer, H. E.; Schwarz, A. D.;
Duchateau, R.; Mountford, P. Chem. Commun. 2010, 46, 273−275.
(1) (a) Drumright, R. E.; Gruber, P. R.; Henton, D. E. Adv. Mater. (13) (a) Liu, X. L.; Shang, X. M.; Tang, T.; Hu, N. H.; Pei, F. K.; Cui,
2000, 12, 1841−1846. (b) Mecking, S. Angew. Chem., Int. Ed. 2004, 43, D. M.; Chen, X. S.; Jing, X. B. Organometallics 2007, 26, 2747−2757.
1078−1085. (c) Williams, C.; Hillmyer, M. Polym. Rev. 2008, 48, 1−
(b) Zhao, W.; Cui, D.; Liu, X.; Chen, X. Macromolecules 2010, 43,
10. (d) Finne, A.; Albertsson, A.-C. Biomacromolecules 2002, 3, 684−
6678−6684.
690. (e) Pang, X.; Zhuang, X.; Tang, Z.; Chen, X. Biotechnol. J. 2010, 5,
(14) (a) Ma, H.; Spaniol, T. P.; Okuda, J. Angew. Chem. 2006, 118,
1125−1136. (f) Li, S.; Liu, D.; Zheng, Y.; Yue, Y.; Huang, Y.; Jing, X.
7982−7985. (b) Ma, H.; Spaniol, T. P.; Okuda, J. Inorg. Chem. 2008,
Acta Polym. Sin. 2012, 9, 1029−1034. (g) Wang, W.; Li, D.; Wang, M.;
47, 3328−3339.
Li, Y.; Gao, C. Chin. J. Polym. Sci. 2011, 29, 233−240.
(15) (a) Hodgson, L. M.; White, A. J. P.; Williams, C. K. J. Polym. Sci.,
(2) (a) Stanford, M. J.; Dove, A. P. Chem. Soc. Rev. 2010, 39, 486−
Part A: Polym. Chem. 2006, 44, 6646−6651. (b) Platel, R. H.;
494. (b) Thomas, C. M. Chem. Soc. Rev. 2010, 39, 165−173.
Hodgson, L. M.; White, A. J. P.; Williams, C. K. Organometallics 2007,
(c) Dijkstra, P. J.; Du, H.; Feijen, J. Polym. Chem. 2011, 2, 520.
26, 4955−4963. (c) Hodgson, L. M.; Platel, R. H.; White, A. J. P.;
(3) (a) Radano, C. P.; Baker, G. L.; Smith, M. R. J. Am. Chem. Soc.
Williams, C. K. Macromolecules 2008, 41, 8603−8607. (d) Platel, R.
2000, 122, 1552−1553. (b) Nomura, N.; Ishii, R.; Akakura, M.; Aoi, K.
J. Am. Chem. Soc. 2002, 124, 5938−5939. (c) Ovitt, T. M.; Coates, G. H.; White, A. J.; Williams, C. K. Inorg. Chem. 2008, 47, 6840−6849.
W. J. Am. Chem. Soc. 2002, 124, 1316−1326. (d) Zhong, Z. Y.; (e) Platel, R. H.; White, A. J.; Williams, C. K. Chem. Commun. 2009,
Dijkstra, P. J.; Feijen, J. Angew. Chem., Int. Ed. 2002, 41, 4510−4513. 27, 4115−4117.
(e) Chen, H.-L.; Dutta, S.; Huang, P.-Y.; Lin, C.-C. Organometallics (16) (a) Cao, T. P.; Buchard, A.; Le Goff, X. F.; Auffrant, A.;
2012, 31, 2016−2025. (f) Tang, Z.; Chen, X.; Pang, X.; Yang, Y.; Williams, C. K. Inorg. Chem. 2012, 51, 2157−2169. (b) Bakewell, C.;
Zhang, X.; Jing, X. Biomacromolecules 2004, 5, 965−970. (g) Zhong, Cao, T. P. A.; Le Goff, X. F.; Long, N. J.; Auffrant, A.; Williams, C. K.
Z.; Dijkstra, P. J.; Feijen, J. J. Am. Chem. Soc. 2003, 125, 11291−11298. Organometallics 2013, 32, 1475−1483.
(h) Majerska, K.; Duda, A. J. Am. Chem. Soc. 2004, 126, 1026−1027. (17) Bakewell, C.; Cao, T. P.; Long, N.; Le Goff, X. F.; Auffrant, A.;
(i) Tang, Z.; Chen, X.; Yang, Y.; Pang, X.; Sun, J.; Zhang, X.; Jing, X. J. Williams, C. K. J. Am. Chem. Soc. 2012, 134, 20577−20580.
Polym. Sci., Part A: Polym. Chem. 2004, 42, 5974−5982. (j) Nomura, (18) Heck, R.; Schulz, E.; Collin, J.; Carpentier, J.-F. J. Mol. Catal. A:
N.; Ishii, R.; Yamamoto, Y.; Kondo, T. Chem.Eur. J. 2007, 13, Chem. 2007, 268, 163−168.
4433−4451. (k) Chisholm, M. H.; Gallucci, J. C.; Quisenberry, K. T.; (19) Arnold, P. L.; Buffet, J. C.; Blaudeck, R. P.; Sujecki, S.; Blake, A.
Zhou, Z. Inorg. Chem. 2008, 47, 2613−2624. (l) Du, H. Z.; Velders, A. J.; Wilson, C. Angew. Chem., Int. Ed. 2008, 47, 6033−6036.
H.; Dijkstra, P. J.; Zhong, Z. Y.; Chen, X. S.; Feijen, J. Macromolecules (20) Otero, A.; Fernandez-Baeza, J.; Lara-Sanchez, A.; Alonso-
2009, 42, 1058−1066. (m) Nomura, N.; Hasegawa, J.; Ishii, R. Moreno, C.; Marquez-Segovia, I.; Sanchez-Barba, L. F.; Rodriguez, A.
Macromolecules 2009, 42, 4907−4909. (n) Spassky, N.; Wisniewski, M. Angew. Chem., Int. Ed. 2009, 48, 2176−2179.
M.; Pluta, C.; LeBorgne, A. Macromol. Chem. Phys. 1996, 197, 2627− (21) Buffet, J.-C.; Kapelski, A.; Okuda, J. Macromolecules 2010, 43,
2637. 10201−10203.
(4) (a) Hormnirun, P.; Marshall, E. L.; Gibson, V. C.; White, A. J. P.; (22) (a) Mazzeo, M.; Tramontano, R.; Lamberti, M.; Pilone, A.;
Williams, D. J. J. Am. Chem. Soc. 2004, 126, 2688−2689. (b) Zhao, W.; Milione, S.; Pellecchia, C. Dalton Trans. 2013, 42, 9338−9351. (b) Li,
Wang, Y.; Liu, X.; Chen, X.; Cui, D.; Chen, E. Y. Chem. Commun. G.; Lamberti, M.; Mazzeo, M.; Pappalardo, D.; Roviello, G.; Pellecchia,
2012, 48, 6375−6377. C. Organometallics 2012, 31, 1180−1188. (c) Platel, R. H.; White, A.
(5) (a) Cheng, M.; Attygalle, A. B.; Lobkovsky, E. B.; Coates, G. W. J. J.; Williams, C. K. Inorg. Chem. 2011, 50, 7718−7728. (d) Grunova, E.;
Am. Chem. Soc. 1999, 121, 11583−11584. (b) Chamberlain, B. M.; Kirillov, E.; Roisnel, T.; Carpentier, J. F. Dalton Trans. 2010, 39,
Cheng, M.; Moore, D. R.; Ovitt, T. M.; Lobkovsky, E. B.; Coates, G. 6739−6752.
W. J. Am. Chem. Soc. 2001, 123, 3229−3238. (c) Chisholm, M. H.; (23) Otero, A.; Fernández-Baeza, J.; Lara-Sánchez, A.; Sánchez-
Gallucci, J.; Phomphrai, K. Inorg. Chem. 2002, 41, 2785−2794. Barba, L. F. Coord. Chem. Rev. 2013, 257, 1806−1868.
(d) Marshall, E. L.; Gibson, V. C.; Rzepa, H. S. J. Am. Chem. Soc. 2005, (24) (a) Fischbach, A.; Klimpel, M. G.; Widenmeyer, M.; Herdtweck,
127, 6048−6051. E.; Scherer, W.; Anwander, R. Angew. Chem., Int. Ed. 2004, 43, 2234−
(6) (a) Cushion, M. G.; Mountford, P. Chem. Commun. 2011, 47, 2239. (b) Meermann, C.; Tornroos, K. W.; Nerdal, W.; Anwander, R.
2276−2278. (b) Chisholm, M. H.; Gallucci, J. C.; Phomphrai, K. Inorg. Angew. Chem., Int. Ed. 2007, 46, 6508−6513. (c) Li, D.; Li, S.; Cui, D.;
Chem. 2004, 43, 6717−6725. Zhang, X. Organometallics 2010, 29, 2186−2193. (d) Evans, W. J.;
(7) (a) Kasperczyk, J. E. Macromolecules 1995, 28, 3937−3939. Giarikos, D. G. Macromolecules 2004, 37, 5130−5132. (e) Fischbach,
(b) Bero, M.; Dobrzynski, P.; Kasperczyk, J. J. Polym. Sci., Part A: A.; Meermann, C.; Eickerling, G.; Scherer, W.; Anwander, R.
Polym. Chem. 1999, 37, 4038−4042. Macromolecules 2006, 39, 6811−6816. (f) Zhiquan, S.; Jun, O.;
(8) (a) Pietrangelo, A.; Hillmyer, M. A.; Tolman, W. B. Chem. Fusong, W.; Zhenya, H.; Fusheng, Y.; Baogong, Q. J. Polym. Sci., Part
Commun. 2009, 45, 2736−2737. (b) Pietrangelo, A.; Knight, S. C.; A: Polym. Chem. 1980, 18, 3345−3357.
Gupta, A. K.; Yao, L. J.; Hillmyer, M. A.; Tolman, W. B. J. Am. Chem. (25) Zhang, Z.; Cui, D. Chem.Eur. J. 2011, 17, 11520−11526.
Soc. 2010, 132, 11649−11657. (26) (a) Garcés, A.; Sánchez-Barba, L. F.; Fernández-Baeza, J.; Otero,
(9) (a) Azor, L.; Bailly, C.; Brelot, L.; Henry, M.; Mobian, P.; A.; Honrado, M.; Lara-Sánchez, A.; Rodríguez, A. M. Inorg. Chem.
Dagorne, S. Inorg. Chem. 2012, 51, 10876−10883. (b) Sauer, A.; 2013, 52, 12691−12701. (b) Cushion, M. G.; Meyer, J.; Heath, A.;
Kapelski, A.; Fliedel, C.; Dagorne, S.; Kol, M.; Okuda, J. Dalton Trans. Schwarz, A. D.; Fernández, I.; Breher, F.; Mountford, P. Organo-
2013, 42, 9007−9023. (c) Chmura, A. J.; Davidson, M. G.; Frankis, C. metallics 2010, 29, 1174−1190. (c) Bigmore, H. R.; Meyer, J.;
J.; Jones, M. D.; Lunn, M. D. Chem. Commun. 2008, 44, 1293−1295. Krummenacher, I.; Ruegger, H.; Clot, E.; Mountford, P.; Breher, F.
(10) (a) Stevels, W. M.; Ankoné, M. J. K.; Dijkstra, P. J.; Feijen, J. Chem.Eur. J. 2008, 14, 5918−5934.
Macromolecules 1996, 29, 3332−3333. (b) Chamberlain, B. M.; Sun, (27) Rong, W.; Liu, D.; Zuo, H.; Pan, Y.; Jian, Z.; Li, S.; Cui, D.
Y.; Hagadorn, J. R.; Hemmesch, E. W.; Young, V. G.; Pink, M.; Organometallics 2012, 32, 1166−1175.
Hillmyer, M. A.; Tolman, W. B. Macromolecules 1999, 32, 2400−2402. (28) Anwander, R.; Runte, O.; Eppinger, J.; Gerstberger, G.;
(11) (a) Cai, C. X.; Amgoune, A.; Lehmann, C. W.; Carpentier, J. F. Herdtweck, E.; Spiegler, M. J. Chem. Soc., Dalton Trans. 1998, 847−
Chem. Commun. 2004, 40, 330−331. (b) Amgoune, A.; Thomas, C. 858.
M.; Roisnel, T.; Carpentier, J.-F. Chem.Eur. J. 2006, 12, 169−179. (29) (a) Herrmann, W. A.; Eppinger, J.; Spiegler, M.; Runte, O.;
(c) Bouyahyi, M.; Ajellal, N.; Kirillov, E.; Thomas, C. M.; Carpentier, Anwander, R. Organometallics 1997, 16, 1813−1815. (b) Eppinger, J.;
J.-F. Chem.Eur. J. 2011, 17, 1872−1883. (d) Amgoune, A.; Thomas, Spiegler, M.; Hieringer, W.; Herrmann, W. A.; Anwander, R. J. Am.
C. M.; Carpentier, J. F. Macromol. Rapid Commun. 2007, 28, 693−697. Chem. Soc. 2000, 122, 3080−3096. (c) Meermann, C.; Gerstberger, G.;

2240 dx.doi.org/10.1021/ma500209t | Macromolecules 2014, 47, 2233−2241


Macromolecules Article

Spiegler, M.; Törnroos, K. W.; Anwander, R. Eur. J. Inorg. Chem. 2008,


12, 2014−2023.
(30) (a) Pan, Y.; Xu, T.; Yang, G.-W.; Jin, K.; Lu, X.-B. Inorg. Chem.
2013, 52, 2802−2808. (b) Gao, W.; Cui, D. J. Am. Chem. Soc. 2008,
130, 4984−4991. (c) Kuo, P.-C.; Chang, J.-C.; Lee, W.-Y.; Lee, H. M.;
Huang, J.-H. J. Organomet. Chem. 2005, 690, 4168−4174.
(31) Sánchez-Barba, L. F.; Garcés, A. S.; Fernández-Baeza, J.; Otero,
A.; Alonso-Moreno, C.; Lara-Sánchez, A. n.; Rodríguez, A. M.
Organometallics 2011, 30, 2775−2789.
(32) (a) Shang, X. M.; Liu, X. L.; Cui, D. M. J. Polym. Sci., Part A:
Polym. Chem. 2007, 45, 5662−5672. (b) Gao, W.; Cui, D.; Liu, X.;
Zhang, Y.; Mu, Y. Organometallics 2008, 27, 5889−5893. (c) Mazzeo,
M.; Lamberti, M.; D’Auria, I.; Milione, S.; Peters, J. C.; Pellecchia, C. J.
Polym. Sci., Part A: Polym. Chem. 2010, 48, 1374−1382.
(33) Ma, H.; Okuda, J. Macromolecules 2005, 38, 2665−2673.
(34) (a) Lappert, M. F.; Pearce, R. J. Chem. Soc., Chem. Commun.
1973, 126−126. (b) Arndt, S.; Voth, P.; Spaniol, T. P.; Okuda, J.
Organometallics 2000, 19, 4690−4700.
(35) Zhang, Z. C.; Cui, D. M.; Trifonov, A. A. Eur. J. Inorg. Chem.
2010, 15, 2861−2866.

2241 dx.doi.org/10.1021/ma500209t | Macromolecules 2014, 47, 2233−2241

You might also like